Oriented Films of Conjugated 2D Covalent Organic Frameworks as

Dec 17, 2017 - ... π-stack in the third dimension to create photoactive porous frameworks. ... available by participants in Crossref's Cited-by Linki...
0 downloads 0 Views 2MB Size
Subscriber access provided by READING UNIV

Article

Oriented Films of Conjugated 2D Covalent Organic Frameworks as Photocathodes for Water Splitting Torben Sick, Alexander Georg Hufnagel, Jonathan Kampmann, Ilina Kondofersky, Mona Calik, Julian M. Rotter, Austin M. Evans, Markus Döblinger, Simon Herbert, Kristina Peters, Daniel Boehm, Paul Knochel, Dana D Medina, Dina Fattakhova-Rohlfing, and Thomas Bein J. Am. Chem. Soc., Just Accepted Manuscript • DOI: 10.1021/jacs.7b06081 • Publication Date (Web): 17 Dec 2017 Downloaded from http://pubs.acs.org on December 17, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Journal of the American Chemical Society is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 9 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

Oriented Films of Conjugated 2D Covalent Organic Frameworks as Photocathodes for Water Splitting Torben Sick‡1, Alexander G. Hufnagel‡1, Jonathan Kampmann‡1, Ilina Kondofersky1, Mona Calik1, Julian M. Rotter1, Austin Evans1, Markus Döblinger1, Simon Herbert1, Kristina Peters1, Daniel Böhm1, Paul Knochel1, Dana D. Medina1, Dina Fattakhova-Rohlfing2,3* and Thomas Bein1* 1

Department of Chemistry and Center for NanoScience (CeNS), University of Munich (LMU), Butenandtstraße 5-13, 81377 Munich, Germany. 2

Forschungszentrum Jülich GmbH, Institute of Energy and Climate Research (IEK-1) Materials Synthesis and Processing, Wilhelm-Johnen-Straße, 52425 Jülich, Germany.

3

Faculty of Engineering and Center for Nanointegration Duisburg-Essen (CENIDE), University of DuisburgEssen, Lotharstraße 1, 47057 Duisburg, Germany ABSTRACT: Light-driven water electrolysis at a semiconductor surface is a promising way to generate hydrogen from sustainable energy sources, but its efficiency is limited by the performance of available photoabsorbers. Here we report the first time investigation of covalent organic frameworks (COFs) as a new class of photoelectrodes. The presented 2DCOF structure is assembled from aromatic amine-functionalized tetraphenylethylene and thiophene dialdehyde building blocks to form conjugated polyimine sheets, which π-stack in the third dimension to create photoactive porous frameworks. Highly oriented COF films absorb light in the visible range to generate photo-excited electrons that diffuse to the surface and are transferred to the electrolyte resulting in proton reduction and hydrogen evolution. The observed photoelectrochemical activity of the 2D-COF films and their photocorrosion stability in water pave the way for a novel class of photoabsorber materials with versatile optical and electronic properties that are tunable through the selection of appropriate building blocks and their three-dimensional stacking.

compensated by nanostructuring.2-4 On the other hand, Si and III-V semiconductors offer favorable charge transport properties and high photocurrents but are prone to photocorrosion unless protected by overlayers.5-6 For this reason, the discovery of new photoabsorber materials is urgent and currently the focus of combinatorial synthetic and computational research.7-10

Introduction Photoelectrochemical (PEC) water splitting is an attractive way to generate hydrogen using renewable energy. Due to the potential of directly converting solar power into a chemical fuel or process feedstock, this process has attracted broad research interest. An ideal photoabsorber will combine efficient light harvesting with suitable band energies for water oxidation and reduction, fast interfacial charge transfer kinetics to the electrolyte, and corrosion stability under operating conditions.1 However, the performance of known systems is modest, being primarily limited by photoabsorber material properties. On the one hand, metal oxides tend to be stable under operation (i.e. resistant against photocorrosion) but exhibit poor semiconductor properties such as short charge carrier diffusion lengths and fast recombination, which can be partially

In this work we explore covalent organic frameworks (COFs) as photoelectrodes for light-driven water splitting. COFs are an emerging class of crystalline polymers composed of organic units linked via covalent bonds to form porous networks. By using multidentate building blocks, two- or three-dimensional frameworks with a defined pore size and high specific surface area in conjunction with appreciable thermal and chemical stability can be obtained. In 2D-COFs, covalently bonded units form two- dimensional (2D) sheets, which stack due to dispersive forces (π-stacking) in the third dimension to

1

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

construct extended open porous frameworks. The πstacking mediates electronic interactions between the functional units, thereby providing another possible path for charge carrier transport in addition to transfer within the covalent sheets. The selection of appropriate building

blocks, linkage motifs (from unconjugated boroxines, boronate esters and borosilicates11-14 to conjugated imines15-19, imides20-21 and others22-24), as well as their stacking mode all provide ways to

Figure 1. a) Synthetic approach for the formation of BDT-ETTA COF with a structural overview of the resulting 2D layers. Due to π-interactions the sheets stack in the third dimension to form the final hexagonal AA eclipsed framework b) TEM image of the resulting powder. c) PXRD of BDT-ETTA (red), comparison to a Pawley-refined pattern (blue) and difference (black line). d) Nitrogen physisorption isotherm of BDT-ETTA with a pore size distribution revealing two distinct pore sizes.

tailor the optical and electronic properties of COF structures, thereby clearing the way for novel materials for optoelectronic systems. COFs have found applications in gas storage25-26, catalysis27-28, separation29-31, energy storage32 and proton conduction33. Zeolitic imidazolate frameworks (ZIFs)34, carbon nitrides (CNs)35-40, phenyl triazine oligomers (PTOs)41, poly-(azomethine) networks (ANWs)42, triazine containing organic frameworks and even covalent organic frameworks (CTFs and COFs)43-50 were found to be active in photocatalysis and photocatalytic hydrogen evolution reactions, where photogenerated charge separation was achieved by adding noble metal co-catalysts and/or sacrificial electron donors. Fabricated as oriented COF thin films, COFs have already found use as active materials in optoelectronic devices51-54. However, COF films being utilized as

photoelectrodes for direct water splitting has yet to be reported. Herein, we demonstrate the first use of an imine-based COF, grown as oriented films on transparent, conducting substrates serving as a light-absorbing material. This is the first report of a COF acting as a photoelectrode to enable photoelectrochemical water splitting in aqueous electrolytes without the use of a cocatalyst or sacrificial agent. In addition, we demonstrate a four-fold increase in the obtained photocurrent by the application of a Pt nanoparticle hydrogen evolution catalyst.

Results and discussion Synthesis of COF systems For the possible use as photocathodes in photoelectrochemical water-splitting devices, we

2

ACS Paragon Plus Environment

Page 2 of 9

Page 3 of 9 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society investigated a 2D-COF built from aromatic amino- and thiophene-based units. The use of nearly planar p-type functional building units linked via polarizable imine bonds is expected to result in conjugated p-type semiconductors, whose optoelectronic properties can be varied via the structure of the building units and the nature of their stacking in the third dimension.

One building unit is a conjugated aromatic and fourfold amine-functionalized tetraphenylethylene (1,1’,2,2’tetra-p-aminophenylethylene, ETTA). The ETTA monomer has been investigated in 2D-imine COF powders55, where it was

Figure 2. Thin (ca. 100 nm) (a, c) and thick (ca. 500 nm) (b, d) BDT-ETTA films grown on ITO substrates: SEM images (crosssection) (a, b) and grazing incidence diffraction (GID) patterns (c, d) revealing a high degree of film orientation.

shown to have a strong impact on geometry, shape, crystallinity and on the stacking distance of adjacent layers. The other component is the linear dialdehyde benzo[1,2-b:4,5-b’]dithiophene-2,6-dicarboxaldehyde (BDT), which is a donor-type dithiophene (Figure 1a).

mesitylene-dioxane (V:V = 1:1) solution in an autoclave. After adding a catalytic amount 6M acetic acid, the films were grown for 72 h at 120 °C. Intensely orange-colored BDT-ETTA COF films were obtained on FTO or ITO substrates. Cross-sectional SEM images of BDT-ETTA films (Figure 2) with different thicknesses demonstrate that COF films can be grown in the thickness range between 100 to 500 nm and are well adhered to the substrate. The surface exhibits some roughness, which is notably less pronounced in the case of thinner films (compare Figure 2a).

For use as electrodes in photoelectrochemical cells, the COFs were grown as films on transparent conducting FTO or ITO substrates. We found that high concentrations of reactants resulted in homogeneous nucleation and growth of bulk COF powders, whereas dilute solutions promote the heterogeneous film growth on a substrate. To avoid precipitation/sedimentation of COF powders on the substrate, the substrates were placed with the ITO side downwards in a PTFE film holder in a

BDT-ETTA COF is well ordered and crystalline, as established by an intense and sharp 100 reflection and the presence of well-defined higher-order reflections in the

3

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

PXRD pattern (Figure 1c). The reflection at around 20° 2θ represents the π-stacking distance, which corresponds to 0.44 nm. The experimentally obtained XRD pattern agrees well with the simulated pattern calculated for an AA eclipsed layer stacking for a hexagonal structure in P6 symmetry (see Figure 1c and SI Figure 13). The π-stacking distance is high in comparison to other COF systems,

which can be explained by the strong contribution of the propeller-shaped (i.e. non-planar) ETTA motif to the final geometry compared to the planar thiophene linker.19 Notably, both the XRD measurements in reflection mode (SI Figure 11) as well as grazing incidence diffraction (GID) patterns (Figure 2c,d) provide strong evidence for

Figure 3. a) Absorbance spectrum of a BDT-ETTA thin film on ITO with a photograph of a representative sample masked with a PFTE adhesive tape (inset). b) Tauc plot analysis of a BDT-ETTA film on ITO showing a direct optical band gap of 2.47 eV. c) Cyclic voltammogram of a BDT-ETTA electrode in non-aqueous electrolyte. d) Calculated alignment between the HOMO and LUMO of BDT-ETTA and the water-splitting redox couples.

oriented growth of the COF films on the substrates. The orientation in c-direction is evident from the intense reflections at q(y)=0 originating from COF layers oriented parallel to the substrate and correspond to the reflections of the BDT-ETTA COF (compare PXRD in Figure 1c). A weak diffuse arc originates from randomly distributed COF particles present on top of the highly-oriented film (see SI Figures 9 and 11). For the electrochemical investigations we used the thinner films as they constitute a well-defined system with less contributions from unordered COF material.

diameter. The pore sizes match the predicted values for the geometry of an AA eclipsed framework. The TEM image of BDT-ETTA (Figure 1b) reveals a high degree of crystallinity and order, recognizable by the large domain sizes. 2D honeycomb-type facets are visible, where the ab plane is oriented perpendicular to the viewing direction. In other viewing directions, channels indicate the growth orientation, highlighting the crystallinity of the COF material with domain sizes of 50100 nm.

Photoelectrochemical properties of COF films The remarkable stability of the obtained BDT-ETTA in different solvents, including water in a pH range from 3 to 14 (see SI Figure 12 for further details), and strong absorption of visible light led us to believe the novel BDTETTA COF structure would be an interesting example of photoabsorber materials for water-splitting applications.

The nitrogen physisorption isotherms of the COF powder demonstrate that the BDT-ETTA COF forms a porous structure with clearly distinguishable micro- and mesopores (Figure 1d). The BET surface areas and total pore volumes for COF bulk material was calculated to be 1360 m2g-1 and 1.0 cm3g-1, respectively. The bimodal pore size distribution shows pores of 1.67 nm and 3.68 nm in

4

ACS Paragon Plus Environment

Page 4 of 9

Page 5 of 9 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society Therefore we measured the photochemical activity at pH 7.

To determine the absolute positions of the conduction and valence band edges of BDT-ETTA films (corresponding to their LUMO and HOMO energies, respectively), electrochemical measurements in a nonaqueous electrolyte (0.1 M NBu4PF6 in anhydrous acetonitrile) were performed. The cyclic voltammogram of a BDT-ETTA film electrode (Figure 3c) shows an anodic peak with an onset at 0.37 V vs. ferrocene/ferrocenium (FOC), which we attribute to the expected oxidation of thiophene moieties. From this, the approximate position of the HOMO can be

The absorption spectrum is shown in Figure 3. BDTETTA films absorb light strongly in the visible range with an absorption threshold of around 550 nm and two absorbance maxima at 360 and 430 nm (Figure 3a). The Tauc plot analysis of BDT-ETTA thin films (Figure 3b) reveals a direct optical band gap of 2.47 eV, which is favorable for photoelectrochemical water splitting.2

Figure 4. a) Linear sweep voltammograms of BDT-ETTA films on ITO performed in the dark (black) and under AM 1.5 illumination through the substrate (red). b) The corresponding IPCE spectrum quantifies the photoresponse of the COF electrodes in the visible spectral range. c) Chronoamperometric data of a BDT-ETTA film collected under chopped illumination 17 -1 -2 (8.3 mHz, 455 nm, 10 s cm ) demonstrating the photocurrent response at different potentials. d) Cyclic voltammograms of BDT-ETTA films grown from different solvents. e) Chronoamperometric data recorded on a BDT-ETTA film at 0.4 V vs. RHE (black) under chopped AM 1.5 illumination. Oxidation current recorded simultaneously on a platinum mesh indicator electrode (red) indicating the formation of hydrogen under illumination (see SI for experimental details).

calculated to be -5.51 eV (see SI for further details). Using the optical bandgap of 2.47 eV determined above, the LUMO position is approximately -3.34 eV. The absolute energy of the HOMO and LUMO of BDT-ETTA as well as the H2O/H2 (hydrogen evolution reaction, HER) and O2/H2O (oxygen evolution reaction, OER) redox couples is plotted in Figure 3d. If we assume that the HOMO and LUMO positions do not change significantly due to protonation or deprotonation of the COF structure, the LUMO is higher in energy than the H2O/H2 redox pair in solution over the entire pH range, which means that photoexcited electrons at the COF surface

should be able to spontaneously transfer to the electrolyte, resulting in proton reduction and hydrogen evolution. The driving force, i.e. the potential difference between the LUMO of the COF and the HER is significant, particularly in acidic solutions. In alkaline solutions, the HOMO energy lies below that of the OER redox couple, which would also render photoelectrochemical water oxidation and thus bias-free one-photon water-splitting thermodynamically possible. However, the potential difference is limited to approximately 0.5 V, which approaches the minimum overpotential needed to drive the OER. Therefore, in the

5

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

absence of an efficient OER catalyst, this reaction is not expected to yield significant currents. The photoelectrochemical performance of 100 nm BDT-ETTA films was characterized by linear sweep voltammetry in nitrogen-purged 0.1 M Na2SO4 aqueous electrolyte under AM 1.5 illumination through the substrate in the potential range between 1.1 and 0.2 V vs. RHE (Figure 4a). The photoelectrode shows an early HER onset potential of 1.0 V vs. RHE reaching currents of up to 1.5 μA cm-2 at 0.2 V vs. RHE. We note that thicker (500 nm) films with a less homogeneous orientation (cf. Figure 2b) show a higher current density of 4.3 μA cm-2 at 0.3 V vs. RHE, demonstrating that the PEC performance of the COF films can be improved (SI Figure 2). We investigated the cause of the photoactivity using electrodes made by dropcasting the COF components BDT and ETTA individually on ITO substrates. Cyclic voltammograms of these are shown in SI figure 3. ETTA shows no difference between dark and light currents below 0.4 V vs. RHE indicating that it is not photoactive by itself. The BDT electrode shows similar but lower photoactivity compared to the COF. We assume, therefore, that the photoactivity of the COF arises from the BDT component and that the formation of an oriented film in which the BDT is covalently bound amplifies its photoresponse and also improves stability. We also investigated whether the oriented porous structure of the BDT-ETTA films is required for successful water reduction. To this end, we prepared reference films using different solvents (mesitylene or anisole) for the COF growth solution, which do not result in oriented growth. Figure 4d shows cyclic voltammograms of these films under dark and illuminated conditions. In the case of films grown from mesitylene, no significant currents are observed in either case. Films grown from anisole exhibit very high dark reduction currents above the HER onset potential, indicating a reaction of the film material. The subsequent CV scans under illumination yield a photocurrent lower than the original dark current over a wide potential range. This indicates that the film is neither photoactive nor stable under operating conditions. From these findings we conclude that oriented BDT-ETTA films, grown using suitable procedures such as those described in this study, are required for stable water photoreduction. Therefore, due to the well-defined geometry of the highly oriented thin (100 nm) films, we have chosen these as a system for further photoelectrochemical study. The incidentphoton-to-current-efficiency (IPCE) of these thin films was determined to examine the photoelectrochemical performance of the COF electrode at different wavelengths (Figure 4b). The BDT-ETTA COF showed light-to-current conversion activity over a broad spectral range below 530 nm, reaching a maximum IPCE of 0.38 % at 355 nm. Stability of the photoabsorber is a main requirement for achieving energy payback via photoelectrochemical hydrogen generation.56 To investigate this, we performed chronoamperometric

measurements under chopped illumination (8.3 mHz, 455 nm, 1017 s-1 cm-2) at different potentials between 0.9 and 0.3 V vs. RHE for 10 min each (Figure 4c). BDT-ETTA films showed a stable photocurrent response over the entire potential range while the absolute values of current density followed the trend of the linear sweep voltammograms. Further, we investigated the stability of the COF photoelectrode for an extended period of time at 0.4 V vs. RHE. At this potential, the BDT-ETTA films showed stable behavior with a negligible dark current density and good photoactivity. Chronoamperometric data was recorded at this potential for 5 hours (Figure 4e, black). The sample was alternatingly kept in the dark and illuminated by AM 1.5 simulated sunlight for 15 minutes at a time. After the initial switch-on transient a steady-state photocurrent density of 0.9 µA cm-2 was reached in each illumination step. To ensure that the resulting photocurrent arises from the water reduction, we have designed a four-electrode setup enabling to monitor continuously the hydrogen evolution during this stability test (see experimental details and SI Figure 4 for further information). A platinum indicator electrode was placed next to the photocathode and polarized at 1.1 V vs. RHE to oxidize dissolved hydrogen in the electrolyte. An oxidation current recorded at this electrode indicates hydrogen evolution in the system although quantification of the hydrogen amount via this method is challenging due to a low collection efficiency of the indicator electrode (SI Figure 6). Prior to measurements the indicator electrode was polarized without illumination until a stable background current of 0.06 µA was reached (SI Figure 5). Illumination of the COF film resulted in a photocurrent detected on the photoelectrode (Figure 4e, black) and a simultaneous rise in the hydrogen oxidation current at the indicator electrode (Figure 4e, red). Switching off the light results in a decay of photocurrent as well as the oxidation current. This behavior is stable and repeatable over the course of the measurement, indicating stability of the material under photoelectrochemical operating conditions. In order to rule out that the oxidative current observed in Figure 4e results from any other species than photoelectrochemically evolved hydrogen, we quantified the product with a hydrogen microsensor (Unisense A/S H2-NPLR) with a hydrogen-selective silicone membrane. At a static potential of the COF film of 0.4 V vs. RHE, a direct correlation between the hydrogen evolution and the illumination of the sample with AM1.5 simulated sunlight is evident (SI Figure 8). The long-term stability demonstrated by the COF photoelectrode shows that no components of the COF material are dissolved by photocorrosion. This demonstrates that a BDT-ETTA covalent organic framework can be used as a stable photocathode for PEC water reduction. The conversion efficiency of photoelectrodes can be enhanced by a number of methods, including the application of a co-

6

ACS Paragon Plus Environment

Page 6 of 9

Page 7 of 9 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society catalyst which facilitates the charge transfer to the electrolyte. We demonstrate that this is viable for COF photocathodes by decorating the COF film with platinum nanoparticles. As shown in Figure 5, the BDT-ETTA/Pt films show a four-fold increase in photocurrent compared to bare BDT-ETTA photocathodes. Therefore, we see the potential to improve the efficiency of COF photoelectrodes by combining them with suitable cocatalysts.

Experimental Section All reagents and solvents were obtained from commercial suppliers and used as received. Benzo(1,2b:4,5-b’)dithiophene (BDT, >98%, TCI), benzyl alcohol (BnOH, anhydrous, Sigma-Aldrich), mesitylene (Mes, anhydrous, Sigma-Aldrich), tetrahydrofuran (THF, extra dry, stabilized, Acros Organics), acetonitrile (SigmaAldrich), ITO glass substrates (Visiontek, 12 Ω/sq). FTOcoated glass substrates (Pilkington, 7 Ω/sq).

General procedure for BDT-ETTA COF films: Under argon, BDT (7.88 mg, 0.032 mmol) and ETTA (76.28 mg, 0.016 mmol) were finely ground, added to a PTFE autoclave, and suspended in a mixture of benzyl alcohol and mesitylene (V/V 1:1, 2000 µL). A glass slide holder, containing ITO or FTO slides, was introduced to the liner. Acetic acid (6 M, 200 µL) was added, the autoclave was sealed and the mixture was kept at 120 °C for 3 days. The resulting orange film was rinsed with anhydrous THF and dried under reduced pressure. COF bulk material that precipitated beneath the film substrate holder was filtered and purified in a Soxhlet extractor for 24 h with anhydrous THF. Structural characterization. 1 H NMR spectra were recorded on Bruker AV 400 and AV 400 TR spectrometers. Proton chemical shifts are expressed in parts per million (δ scale) and are calibrated using residual non deuterated solvent peaks as an internal reference (e.g. DMSO-d6: 2.50 ppm). Ultraviolet–vis– infrared diffuse reflectance spectra (Kubelka–Munk spectrum) of the films were recorded on a Perkin-Elmer Lambda 1050 spectrometer equipped with a 150 mm integrating sphere. Thin film absorbance spectra were measured by illuminating the sample though the substrate side and correcting it with the relating reflection.57 Scanning electron microscopy (SEM) images were recorded with a JEOL 6500F and a FEI Helios NanoLab G3 UC scanning electron microscope equipped with a field emission gun operated at 3-5 kV. Transmission electron microscopy (TEM) was performed on an FEI Titan Themis equipped with a field emission gun operated at 300 kV. X-ray diffraction (XRD) measurements were performed using a Bruker D8 Discover with Ni-filtered Cu Kα radiation and a LynxEye position-sensitive detector. Experimental XRD data were used for Pawley refinement to optimize the hypothetical structure. The initial structure models of the COFs were built using the Forcite module of the Accelrys Materials Studio software package. We applied the space group with the highest possible symmetry, i.e. P6, taking into account the propeller-like conformation of the central building blocks. Using this coarse model we determined the unit cell parameters via Pawley refinement of the PXRD data. Nitrogen sorption isotherms were recorded on a Quantachrome Autosorb 1 at 77 K within a pressure range from p/p0 = 0.001 to 0.98. Prior to the measurement of the sorption isotherms the samples were heated for 24

Figure 5. Linear sweep voltammograms of BDT-ETTA films on ITO performed in the dark (black) and under AM 1.5 illumination through the substrate (red). The combination of BDT-ETTA with platinum nanoparticles (solid lines) leads to an increased photocurrent over the whole potential range compared to bare BDT-ETTA films (dashed lines).

Conclusion Our results show that BDT-ETTA COF films are viable photocathodes for light-driven water reduction. The material meets the requirements for this application, which are efficient light harvesting, suitable band positions, and stability. The polyimine framework BDTETTA is the first to be investigated in this context, representing an unexplored field of applications for COFs. Given the enormous diversity of molecular building units that can be employed in the construction of COFs, both the development of new framework structures and finetuning the properties of existing ones is possible. We envision that future research on combinations of building blocks and stacking modes and the combination of the photoabsorber with suitable co-catalysts will allow for further optimizing charge carrier lifetimes, long term stability, and charge transfer kinetics in order to improve the conversion efficiency obtainable from COF-based photoelectrodes.

7

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(9) Singh, A. K.; Mathew, K.; Zhuang, H. L.; Hennig, R. G., J. Phys. Chem. Lett. 2015, 6, 1087-1098. (10) Castelli, I. E.; Olsen, T.; Datta, S.; Landis, D. D.; Dahl, S.; Thygesen, K. S.; Jacobsen, K. W., Energy Environ. Sci. 2012, 5, 58145819. (11) Hunt, J. R.; Doonan, C. J.; LeVangie, J. D.; Côté, A. P.; Yaghi, O. M., J. Am. Chem. Soc. 2008, 130, 11872-11873. (12) Côte, A. P.; Benin, A. I.; Ockwig, N. W.; O'Keeffe, M.; Matzger, A. J.; Yaghi, O. M., Science 2005, 310, 1166-1170. (13) Calik, M.; Sick, T.; Dogru, M.; Döblinger, M.; Datz, S.; Budde, H.; Hartschuh, A.; Auras, F.; Bein, T., J. Am. Chem. Soc. 2016, 138, 12341239. (14) Lohse, M. S.; Rotter, J. M.; Margraf, J. T.; Werner, V.; Becker, M.; Herbert, S.; Knochel, P.; Clark, T.; Bein, T.; Medina, D. D., CrystEngComm 2016, 18, 4295-4302. (15) Medina, D. D.; Rotter, J. M.; Hu, Y.; Dogru, M.; Werner, V.; Auras, F.; Markiewicz, J. T.; Knochel, P.; Bein, T., J. Am. Chem. Soc. 2015, 137, 1016-1019. (16) Uribe-Romo, F. J.; Hunt, J. R.; Furukawa, H.; Klöck, C.; O’Keeffe, M.; Yaghi, O. M., J. Am. Chem. Soc. 2009, 131, 4570-4571. (17) Chen, X.; Addicoat, M.; Irle, S.; Nagai, A.; Jiang, D., J. Am. Chem. Soc. 2013, 135, 546-549. (18) Zhang, Y.-B.; Su, J.; Furukawa, H.; Yun, Y.; Gándara, F.; Duong, A.; Zou, X.; Yaghi, O. M., J. Am. Chem. Soc. 2013, 135, 16336-16339. (19) Ascherl, L.; Sick, T.; Margraf, J. T.; Lapidus, S. H.; Calik, M.; Hettstedt, C.; Karaghiosoff, K.; Döblinger, M.; Clark, T.; Chapman, K. W.; Auras, F.; Bein, T., Nat. Chem. 2016, 8, 310-316. (20) Fang, Q.; Zhuang, Z.; Gu, S.; Kaspar, R. B.; Zheng, J.; Wang, J.; Qiu, S.; Yan, Y., Nat. Commun. 2014, 5, 4503-4510. (21) Fang, Q.; Wang, J.; Gu, S.; Kaspar, R. B.; Zhuang, Z.; Zheng, J.; Guo, H.; Qiu, S.; Yan, Y., J. Am. Chem. Soc. 2015, 137, 8352-8355. (22) Uribe-Romo, F. J.; Doonan, C. J.; Furukawa, H.; Oisaki, K.; Yaghi, O. M., J. Am. Chem. Soc. 2011, 133, 11478-11481. (23) Dalapati, S.; Jin, S.; Gao, J.; Xu, Y.; Nagai, A.; Jiang, D., J. Am. Chem. Soc. 2013, 135, 17310-17313. (24) Jackson, K. T.; Reich, T. E.; El-Kaderi, H. M., Chem. Commun. 2012, 48, 8823-8825. (25) Furukawa, H.; Yaghi, O. M., J. Am. Chem. Soc. 2009, 131, 88758883. (26) Doonan, C. J.; Tranchemontagne, D. J.; Glover, T. G.; Hunt, J. R.; Yaghi, O. M., Nat. Chem. 2010, 2, 235-238. (27) Xu, H.; Chen, X.; Gao, J.; Lin, J.; Addicoat, M.; Irle, S.; Jiang, D., Chem. Commun. 2014, 50, 1292-1294. (28) Ding, S.-Y.; Gao, J.; Wang, Q.; Zhang, Y.; Song, W.-G.; Su, C.-Y.; Wang, W., J. Am. Chem. Soc. 2011, 133, 19816-19822. (29) Oh, H.; Kalidindi, S. B.; Um, Y.; Bureekaew, S.; Schmid, R.; Fischer, R. A.; Hirscher, M., Angew. Chem. Int. Ed. 2013, 52, 1321913222. (30) Ma, H.; Ren, H.; Meng, S.; Yan, Z.; Zhao, H.; Sun, F.; Zhu, G., Chem. Commun. 2013, 49, 9773-9775. (31) Lohse, M. S.; Stassin, T.; Naudin, G.; Wuttke, S.; Ameloot, R.; De Vos, D.; Medina, D. D.; Bein, T., Chem. Mater. 2016, 28, 626-631. (32) DeBlase, C. R.; Silberstein, K. E.; Truong, T.-T.; Abruña, H. D.; Dichtel, W. R., J. Am. Chem. Soc. 2013, 135, 16821-16824. (33) Chandra, S.; Kundu, T.; Kandambeth, S.; BabaRao, R.; Marathe, Y.; Kunjir, S. M.; Banerjee, R., J. Am. Chem. Soc. 2014, 136, 6570-6573. (34) Flugel, E. A.; Lau, V. W.; Schlomberg, H.; Glaum, R.; Lotsch, B. V., Chem. Eur. J. 2016, 22, 3676-3680. (35) Schwinghammer, K.; Tuffy, B.; Mesch, M. B.; Wirnhier, E.; Martineau, C.; Taulelle, F.; Schnick, W.; Senker, J.; Lotsch, B. V., Angew. Chem. Int. Ed. 2013, 52, 2435-2439. (36) Schwinghammer, K.; Mesch, M. B.; Duppel, V.; Ziegler, C.; Senker, J.; Lotsch, B. V., J. Am. Chem. Soc. 2014, 136, 1730-1733. (37) Lau, V. W.; Mesch, M. B.; Duppel, V.; Blum, V.; Senker, J.; Lotsch, B. V., J. Am. Chem. Soc. 2015, 137, 1064-1072. (38) Caputo, C. A.; Gross, M. A.; Lau, V. W.; Cavazza, C.; Lotsch, B. V.; Reisner, E., Angew. Chem. Int. Ed. 2014, 53, 11538-11542.

h at 120°C under turbo-pumped vacuum. For the evaluation of the surface area the BET model was applied between 0.05 and 0.2 p/p0. Pore size distributions were calculated using the QSDFT equilibrium model with a carbon kernel for cylindrical pores.

ASSOCIATED CONTENT The Supporting Information is available free of charge on the ACS Publications website at DOI: XXXXXX Detailed information about synthetic procedures, structural and electrochemical characterization of BDT-ETTA photoelectrodes.

AUTHOR INFORMATION Corresponding Author *[email protected] *[email protected]

Author Contributions T.S., A.G.H, J.K., D.D.M and D.F.-R. conceived and designed the project. T.S., I.K., J.M.R., A.E., M.C., S.H. and D.D.M. carried out the syntheses and characterized the materials. K.P. carried out the SEM characterization. M.D. carried out the TEM characterization. J.K., A.G.H., D.B. and I.K. carried out the electrochemical characterization. T.S., A.G.H. and J.K. wrote the manuscript with contributions of all the authors. D.F-R. and T.B. supervised the project. All authors discussed the results and contributed to the manuscript.

Notes The authors declare no competing financial interest.

ACKNOWLEDGMENTS The authors are grateful for funding from the German Science Foundation (DFG, SPP 1613), the NIM cluster (DFG) and the Free State of Bavaria (the research networks ‘Solar Technologies Go Hybrid’ and UMWELTnanoTECH). The research leading to these results received funding from the European Research Council under the European Union’s Seventh Framework Programme (FP7/2007-2013)/ERC Grant Agreement 321339. A.G.H. gratefully acknowledges funding by the Fonds der chemischen Industrie. A.E. thanks the DAAD RISE worldwide program for the opportunity to conduct this research at the LMU.

REFERENCES (1) Peter, L. M., J. Solid State Electrochem. 2012, 17, 315-326. (2) van de Krol, R.; Liang, Y.; Schoonman, J., J. Mater. Chem. 2008, 18, 2311-2320. (3) Sivula, K.; Le Formal, F.; Gratzel, M., ChemSusChem 2011, 4, 432449. (4) Dunn, H. K.; Feckl, J. M.; Muller, A.; Fattakhova-Rohlfing, D.; Morehead, S. G.; Roos, J.; Peter, L. M.; Scheu, C.; Bein, T., Phys. Chem. Chem. Phys. 2014, 16, 24610-24620. (5) Hu, S.; Shaner, M. R.; Beardslee, J. A.; Lichterman, M.; Brunschwig, B. S.; Lewis, N. S., Science 2014, 344, 1005-1009. (6) Reece, S. Y.; Hamel, J. A.; Sung, K.; Jarvi, T. D.; Esswein, A. J.; Pijpers, J. J.; Nocera, D. G., Science 2011, 334, 645-648. (7) Woodhouse, M.; Parkinson, B. A., Chem. Soc. Rev. 2009, 38, 197210. (8) Woodhouse, M.; Parkinson, B. A., Chem. Mater. 2008, 20, 24952502.

8

ACS Paragon Plus Environment

Page 8 of 9

Page 9 of 9 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society (39) Zhang, J.; Chen, X.; Takanabe, K.; Maeda, K.; Domen, K.; Epping, J. D.; Fu, X.; Antonietti, M.; Wang, X., Angew. Chem. Int. Ed. Engl. 2010, 49, 441-444. (40) Wang, X.; Maeda, K.; Thomas, A.; Takanabe, K.; Xin, G.; Carlsson, J. M.; Domen, K.; Antonietti, M., Nat. Mater. 2009, 8, 7680. (41) Schwinghammer, K.; Hug, S.; Mesch, M. B.; Senker, J.; Lotsch, B. V., Energy Environ. Sci. 2015, 8, 3345-3353. (42) Schwab, M. G.; Hamburger, M.; Feng, X.; Shu, J.; Spiess, H. W.; Wang, X.; Antonietti, M.; Mullen, K., Chem. Commun. 2010, 46, 8932-8934. (43) Vyas, V. S.; Haase, F.; Stegbauer, L.; Savasci, G.; Podjaski, F.; Ochsenfeld, C.; Lotsch, B. V., Nat. Commun. 2015, 6, 8508-8516. (44) Stegbauer, L.; Schwinghammer, K.; Lotsch, B. V., Chem. Sci. 2014, 5, 2789-2793. (45) Bi, J.; Fang, W.; Li, L.; Wang, J.; Liang, S.; He, Y.; Liu, M.; Wu, L., Macromol. Rapid Commun. 2015, 36, 1799-1805. (46) Kuecken, S.; Acharjya, A.; Zhi, L.; Schwarze, M.; Schomacker, R.; Thomas, A., Chem Commun (Camb) 2017, 53, 5854-5857. (47) Chen, X.; Addicoat, M.; Jin, E.; Zhai, L.; Xu, H.; Huang, N.; Guo, Z.; Liu, L.; Irle, S.; Jiang, D., J. Am. Chem. Soc. 2015, 137, 3241-3247. (48) Haase, F.; Banerjee, T.; Savasci, G.; Ochsenfeld, C.; Lotsch, B. V., Faraday Discuss. 2017, 201, 247-264.

(49) He, S.; Rong, Q.; Niu, H.; Cai, Y., Chem. Commun. 2017, 53, 9636-9639. (50) Banerjee, T.; Haase, F.; Savasci, G.; Gottschling, K.; Ochsenfeld, C.; Lotsch, B. V., J. Am. Chem. Soc. 2017. (51) Dogru, M.; Handloser, M.; Auras, F.; Kunz, T.; Medina, D.; Hartschuh, A.; Knochel, P.; Bein, T., Angew. Chem. Int. Ed. 2013, 52, 2920-2924. (52) Calik, M.; Auras, F.; Salonen, L. M.; Bader, K.; Grill, I.; Handloser, M.; Medina, D. D.; Dogru, M.; Löbermann, F.; Trauner, D.; Hartschuh, A.; Bein, T., J. Am. Chem. Soc. 2014, 136, 17802-17807. (53) Medina, D. D.; Werner, V.; Auras, F.; Tautz, R.; Dogru, M.; Schuster, J.; Linke, S.; Döblinger, M.; Feldmann, J.; Knochel, P.; Bein, T., ACS Nano 2014, 8, 4042-4052. (54) Jin, S.; Ding, X.; Feng, X.; Supur, M.; Furukawa, K.; Takahashi, S.; Addicoat, M.; El-Khouly, M. E.; Nakamura, T.; Irle, S.; Fukuzumi, S.; Nagai, A.; Jiang, D., Angew. Chem. Int. Ed. 2013, 52, 2017-2021. (55) Zhou, T. Y.; Xu, S. Q.; Wen, Q.; Pang, Z. F.; Zhao, X., J. Am. Chem. Soc. 2014, 136, 15885-15888. (56) Sathre, R.; Scown, C. D.; Morrow, W. R.; Stevens, J. C.; Sharp, I. D.; Ager, J. W.; Walczak, K.; Houle, F. A.; Greenblatt, J. B., Energy Environ. Sci. 2014, 7, 3264-3278. (57) Klahr, B. M.; Martinson, A. B. F.; Hamann, T. W., Langmuir 2011, 27, 461-468.

For Table of Contents Only

9

ACS Paragon Plus Environment